Influence of Physical and Structural Aspects of Food On Starch Digestion

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Influence of 

Physical and Structural
Aspects of Food on Starch Digestion

Ingrid Contardo and Pedro Bouchon

1  Mass Transfer and Starch Digestion Kinetics

High consumption of foods rich in starch can be linked with some conditions
and diseases such as obesity and type II diabetes (Zimmet, Alberti, & Shaw, 2001).
The physical state of starch granules influences how they are digested while simple
sugars are absorbed in the gastrointestinal system. Even though the field of food
design is seeking to generate starchy foods with beneficial properties for health,
focusing on reducing or slowing the absorption of glucose in the blood, it is impor-
tant to understand how starch is hydrolyzed and absorbed in the body, and how the
design of novel foods can provide them with functionality through their structure
and disintegration.

1.1  General Aspects and Approaches to Modeling

The kinetics of starch digestion are influenced by the physical and chemical charac-
teristics of the starch granules and the interactions with physiological events occur-
ring within the gastrointestinal tract (Bjorck, Granfeldt, Liljeberg, Tovar, & Asp,
1994). The features of starch are also influenced by food composition and structure,
and food processing conditions. Extrinsic factors  that  influence  starch  digestion
include the nature of starch; its physical form; interactions with proteins, lipids, or
sugars; the presence of enzyme inhibitors; food processing; food structure (initial
hardness or porosity); and bolus hydration/disintegration, while intrinsic factors

I. Contardo · P. Bouchon (*)


Department of Chemical and Bioprocess Engineering, Pontificia Universidad Católica de
Chile, Santiago, Chile
e-mail: pbouchon@ing.puc.cl

© Springer Nature Switzerland AG 2019 303


O. Gouseti et al. (eds.), Interdisciplinary Approaches to Food Digestion,
https://doi.org/10.1007/978-3-030-03901-1_15
304 I. Contardo and P. Bouchon

Fig. 1  Extrinsic and intrinsic factors that influence the kinetics of starch digestion [based on
Bornhorst and Singh (2013), Englyst, Englyst, Hudson, Cole, and Cummings (1999), and Guyton
and Hall (2006)]

consist of particle disintegration, activity/concentration/flow rate/viscosity of secre-


tions, mastication time, and mixing (peristalsis and segmentation). These factors are
represented schematically in Fig. 1. In addition, the amount of glucose available for
absorption is a critical factor in the reduction of the blood sugar level, thus starch
hydrolysis could be limited or modulated in order to obtain the desired effects on
luminal glucose absorption.
Static in vitro models have been established to study starch digestibility, which
mimic some physiological conditions for digestion, including physical transforma-
tions and enzymatic hydrolysis, as well as measuring the amount of glucose released
from a food during timed incubation under well determined conditions. Static mod-
els provide technical simplification under rapid, simple, and controlled conditions;
and they allow for preliminary results to be obtained. Englyst, Kingman, and
Cummings (1992) developed an enzymatic in vitro method to classify starchy foods
based on their digestibility for nutritional purposes, where rapidly available glucose
(RAG) was defined as the fraction of glucose  that was obtained after 20  min of
hydrolysis. Slowly available glucose (SAG) was said to be the fraction obtained
between 20 and 120 min of hydrolysis. These amounts of glucose are likely to be
available for rapid and slow absorption, respectively, in the human small intestine.
Finally, unavailable glucose (UG) was defined as the fraction of glucose that is pres-
ent in the food but could not be released after 120 min of hydrolysis. Furthermore,
Goñi, Garcia-Alonso, and Saura-Calixto (1997) developed a static in vitro proce-
dure to measure the rate of starch digestion in starchy common foodstuffs. The pos-
sible estimation of glycemic index (GI) from the percentage of total hydrolyzed
starch, depending on the sampling time, was also analyzed. The authors concluded
that this in vitro procedure could be useful in the estimation of GI. More recently, a
standardized static method was proposed by the COST action INFOGEST network
(Minekus et al., 2014) to harmonize the variety of in vitro protocols found in the
literature as variations in factors such as the amount and type of enzymes, incuba-
tion time, pH, and specified temperature generate different results from the same
Influence of Physical and Structural Aspects of Food on Starch Digestion 305

foods. The proposal involves the simulation of oral, gastric, and intestinal digestion;
and includes a standardized assay for enzymatic activity determination of the
enzymes that are added at each step.
During testing using  static models, the oral, gastric, or intestinal phases are
reproduced by a mono-compartmental step and the specific test conditions
are adapted for each specific application. Typically, in experiments with static mod-
els, the foods are mixed with simulated fluids during a pre-determined  time of
digestion while controlling the environment in temperature and pH and fixing the
concentration of enzymes. Static models lack the simulation of realistic enzyme to
substrate ratios, continued changes in pH, transit times or removal of digested prod-
ucts, in time and place. In contrast, dynamic in vitro models involve increased com-
plexity, with the addition of digestive fluids with control of flow rate and composition
of their secretions, realistic  gastrointestinal profiles of pH, complex peristaltic
movements or mixing segmentation, diffusion, and gastric emptying cycles. One
dynamic model is the TNO Gastrointestinal Model (TIM), developed at The
Netherlands Organization for Applied Scientific Research (TNO) of the Nutrition
and Food Research Centre. This is a multicompartmental model designed to realisti-
cally simulate conditions in the lumen of the gastrointestinal tract. TIM combines a
reproducible and accurate simulation of digestion processes with detailed kinetic
elements. This model has been developed to study the fraction of a compound that
is available for absorption through the gut wall. Under this approach, the model has
been used to predict the glycemic response after intake of carbohydrates (Nalin
et al., 2015). The TIM-Carbo technology was validated against 21 different in vivo
plasma glucose response curves after the intake of carbohydrate food products
(R = 0.91). Another model is the Dynamic Gastric Model (DGM) developed by the
Institute of Food Research (IFR) (Norwich, UK), which can simulate both the bio-
chemical and mechanical aspects of gastric digestion (stomach and antrum) in a
realistic, time-dependent manner. Secretion rates can be adapted dynamically to the
changing conditions of acidification or fill state and the system allows for the use of
complex food matrices (comparable to in vivo studies) and emulates the peristaltic
movements of the stomach in amplitude, intensity, and frequency (Wickham,
Faulks, Mann, & Mandalari, 2012). Similarly, the Human Gastric Simulation (HGS)
is another dynamic in vitro stomach model, which consists of a latex vessel and a
series of rollers secured on belts that are driven by motor and pulleys to create a
continuous contraction of the latex wall in order to simulate the peristaltic move-
ments of stomach walls, with similar amplitude and frequency of contraction forces
as reported in vivo. It also incorporates gastric secretions, emptying systems, and
temperature control that enable accurate simulation of dynamic digestion processes
for detailed investigation of the changes in the physical chemical properties of
ingested foods. Kong and Singh (2010) used HGS during the digestion of rice. They
demonstrated that the amount of acid became a limiting factor for the acid hydroly-
sis as the solids content of emptied digesta was affected by the amount of rice avail-
able for digestion relative to the amount of acid present (see Fig. 2).
Some differences between in  vivo and in  vitro methods used to study starch
digestibility from solid foods exist, as in vitro methods cannot emulate all the com-
306 I. Contardo and P. Bouchon

Fig. 2  Profile of pH and change of solids content in emptied digesta during in vitro digestion of
rice using Human Gastric Simulation (HGS) (mean of n  =  3) [extracted from Kong and Singh
(2010)]

plex interactions among extrinsic (food characteristics) and intrinsic (body


­physiological) aspects involved in the digestion process (e.g., the nature of the expo-
sure to salivary α-amylase prior to simulated gastric phase digestion, simultaneous
peristaltic movements, gradual pH changes of the food digesta, or origin of
α-amylase). Generally, in vitro approaches to quantify starch digestibility are per-
formed in a two-stage (gastric digestion and intestinal absorption) simulated diges-
tion system. However, the human digestive tract is a complex system of organs
which can be segmented into at least four main units: the mouth, the stomach, the
small intestine, and the large intestine. Starch is affected in different ways in these
steps. Figure 3 shows a representative summary of hydrolysis of starch and absorp-
tion of glucose in the gastrointestinal tract focusing on the main processes (into the
mouth, stomach, small intestine, and large intestine) and highlighting the key vari-
ables that may influence starch digestion kinetics.
Various mathematical models have been used to study the kinetics of starch
digestion after the ingestion of solid starchy foods. As discussed, starch digestion is
a complex process involving a simultaneous mass transfer related to bolus forma-
tion, bolus disintegration, hydrolysis of starch, and glucose release and absorption.
Likewise, digestibility of starch-containing solid foods is widely affected by the
physical–chemical characteristics of starch granules and interactions with the food
matrix (component interactions or food structure). The kinetics of starch digestion
frequently show simple decay curves with apparent first-order behavior; this can
include the rate of starch degradation, the rate of glucose release, or the appearance
Influence of Physical and Structural Aspects of Food on Starch Digestion 307

Organ/unit Physical aspects Chemical aspects


(secretions) Process Variables Process Variables
(function) (function)
Mouth Mastication Mastication time Enzymatic hydrolysis Salivary secretions
(salivary α- (milling, lubrication, Mixing time (initial starch breakdown by flow rate/mass
amylases, mixing, bolus Salivary secretions flow salivary α-amylase) Activity of
water, mucin, formation) rate/mass enzyme
salts) Hardness/porosity of Output: starch, maltose, Residence time
solid food maltotriose, and α-limit
Particle hydration dextrin
Output: bolus Viscosity
Stomach Peristalsis and mixing Degree of mixing Enzymatic hydrolysis Gastric secretions
(acid, gastric waves Secretions flow rate/mass Protein and lipid breakdown flow rate/mass
enzymes, (bolus transport, Bolus volume Acid hydrolysis Activity of
water, salt, particle disintegration, Particle size pH regulation (pH>4 to enzymes
mucus) mixing, pyloric pump) Rate of bolus amylase) Residence time
Gastric emptying disintegration Starch breakdown Enzyme inhibitors
(propel of chyme) Viscosity pH profile
Output: starch, maltose,
maltotriose, and α-limit
Output: Chyme dextrin
Small Peristalsis Chyme flow rate/mass Enzymatic hydrolysis Activity/secretion
intestine (chyme transport) Residence time (main starch, maltose, of enzymes
(alkaline Segmentation Adsorption of enzymes maltotriose, and α-limit Residence time
mucus, (mixing to facilitate onto solid substrates dextrin breakdown by Secretions flow
enterocytes, absorption) Available glucose pancreatic α-amylases) rate/mass
pancreatic α- Convection - Viscosity pH regulation Enzyme inhibitors
amylases, Diffusion Digestible starch
water, bile, (active/passive) pH profile
salt) (absorption of glucose)

Output: absorption of Output: glucose


glucose to bloodstream
Large Haustrations Degree of mixing Fermentation Microbial
intestine (mixing to facilitate Particle size (substrate) population
(alkaline fermentation) Flow rate Bacterial growth
mucus, Mass movements Pressure Output: production of short- Concentration of
bicarbonate (Resistant starch chain fatty acids (SCFA) resistant starch
ions) transport) (acetate, propionate,
butyrate), oligosaccharides,
Output: absorption of and lactate
vitamins and water

Fig. 3  Representative overview of the processes during starch hydrolysis and absorption in the
gastrointestinal tract (mouth, stomach, small intestine, and large intestine), highlighting physical
and chemical aspects influencing starch digestion kinetics from solid food, and the key variables in
an engineering approach [adapted from Bornhorst and Singh (2014)]

of oligosaccharides over time. Thus, empirical exponential models (e.g., first-order


equation and Weibull function) and Michaelis–Menten models have been used to fit
the kinetics of starch hydrolysis or glucose release from starch. Accordingly, using
empirical exponential models, the end-point product concentration, the pseudo first-­
order digestibility rate constant and the fraction of digested starch can be fitted by
linear fits to logarithmic plots (Log of Slope (LOS)); through which it has been
shown that starch amylolysis in solid foods occurred via a two-phase system (rapid
and slower phase) reflecting differences in substrate accessibility; also, glucose was
found to inhibit amyloglucosidase after a “grace” period at low glucose concentra-
tions; and small starch particle sizes increase the rate of enzymatic action.
Likewise, amylolysis progress-curves can be well described by Weibull function
(Dona, Pages, Gilbert, & Kuchel, 2010). For instance, Goñi et al. (1997) used a first-­
order equation that governs the hydrolytic process:


(
C = C∞ l − e − kt ) (1)
308 I. Contardo and P. Bouchon

100
White bread Spaghetti Rice Biscuits
Lentils Chick peas Beans Peas
Boiled potatoes Crisp potatoes
80
Total starch hydrolysis (%)

60

40

20

0
0 30 60 90 120 150 180
Time (min)

Fig. 4  Total starch hydrolysis rate, using a first-order equation [extracted from Goñi et al. (1997)]

Where C is the concentration of hydrolyzed starch (kg/kg) at t time (min), C∞ is the


equilibrium concentration (concentration of digestible starch in the stage where no
more product is formed), k is the kinetic constant (min−1) and t is the chosen time
(min). Experimental values for each investigated starchy food were adjusted to the
equation to calculate k and C∞, using LOS plots. In general, the k values reflect the
susceptibility of starch or starchy foods towards hydrolysis by amylase, and typi-
cally range from 10−5 to 10−3 min−1 but are dependent on enzyme concentration. The
decay equation suggests that the rate of reaction decreases with time due to sub-
strate depletion and it is related to the diffusion of amylase onto the surface of the
granule particle and the physical nature of substrate. The curves of starch digestion
show a first part, where the starch hydrolysis rate increases, and a second, where a
maximal plateau level is slowly reached (see Fig. 4). Additionally, the hydrolysis
indices (HI) from the determination of the area under the curves (AUC) obtained by
integration of the first order rate equation is a simple alternative to the glycemic
index (GI) values for starchy foods. Moreover, amylolysis progress-curves can be
well described by Weibull function (Dona et al., 2010). Likewise, Wang, Zeng, Liu,
and Yuan (2006) used an exponential model considering taking into account product
inhibition to study the hydrolysis of corn starch by glucoamylase:

n
ds  ds   p 
=   exp ( − ki ( p − p0 ) )  1 −  (2)
dt  dt  p =0  pm 

Influence of Physical and Structural Aspects of Food on Starch Digestion 309

Where S is substrate concentration (kg · L−1), t is time (h), ki is the product inhibi-


tion constant (L · kg−1), p is product concentration (kg · L−1), p0 is the minimum
product concentration (kg · L−1) below which product inhibition is not observed, pm
is the product concentration obtained when all the starch in the system is hydrolyzed
into sugars (kg · L−1), and n is an empirical constant (non-dimensional). The experi-
ments showed that a product inhibition effect appeared when product concentration
was above p0.
Furthermore, another empirical model used to explain starch digestion is the
Michaelis–Menten model, which describes reactions where the enzyme concentra-
tion is small relative to the substrate concentration:

Vmax S
v= (3)
Km + S

Where v is the reaction velocity (kg · L  · min−1), vmax is the maximum reaction
−1

velocity (kg · L−1 · min−1), Km is the Michaelis–Menten constant which gives an idea


of the affinity between enzyme and substrate and represents the concentration of
digestible (available) starch (kg · L−1), and S is total starch concentration (kg · L−1).
In addition, vmax = kcat*E0, where kcat is the catalytic constant for the enzyme and E0
is the total enzyme concentration (kg · L−1); kcat involves the number of molecules of
product formed per molecule of enzyme in unit time (min−1). This model can be
employed to describe the release of glucose in the initial stages (within the first
10–15 min) as a function of starch concentration before product inhibition and sub-
strate exhaustion (Dona et al., 2010). In hydrothermally processed starches (e.g.,
boiling process at 100 °C), the reduction in Km occurs as a result of the increased
substrate availability due to the loss of the semi-crystalline native starch granule
structure. Additionally, the gelatinization process affects the catalytic efficiency
(kcat/Km), increasing its values after the hydrothermal process (e.g., values increased
by 13-fold (waxy rice) to 239-fold (potato) after the boiling process). Similarly, this
phenomenon increases the kcat values in comparison to native starch (Slaughter,
Butterworth, & Ellis, 2001). The modifications of this model are related to devia-
tions from Michaelis–Menten kinetics in the early stages of hydrolysis of starch
affected by the starch structure (e.g., the degree to which amylose leaching occurs
during gelatinization), the complexity of the food (bioaccessibility of starch), or
when product inhibition and substrate exhaustion start to happen. Thus, the
competitive-­inhibition model from the Michaelis–Menten framework has been used
to describe product/substrate inhibition of glucoamylase, incorporating high coef-
ficients of variation on parameter estimates. An increase in starch concentration
involved in  the digestion profile means that the system  no longer adheres to
Michaelis–Menten behavior. Also, an inhibitory effect from this type of substrate,
which may cause differences in the mass transfer of the molecules (e.g., by a higher
apparent viscosity, and different rheological or structural features of starch), may
further increase mass-transfer resistance. Under this approach, the use of the linear-
ized graphical method such  as Lineweaver–Burk plots obtained from inhibition
experiments has not been able to distinguish between the two mechanisms, promot-
ing the use of multiple models to describe the kinetics of starch hydrolysis.
310 I. Contardo and P. Bouchon

1.2  Oral Processing

Breakdown behavior in the mouth and stomach of solid starchy products can deter-
mine the accessibility of starch granules during the hydrolysis process. In the mouth,
the physical and chemical transformations of starch-based foods by simultaneous
processes take place; food mastication, lubrication, mixing, and bolus formation.
The main role of mastication is the reduction of food in terms of  particle size.
Usually, breakdown behavior is described in terms of the number of particles of dif-
ferent sizes present in the chewed/digested food, typically characterized by bimodal
behavior. For instance, such data are used to fit the cumulative distribution of par-
ticles in terms of either their numbers or surface areas to a Rosin–Rammler model
(unimodal Weibull distribution), and quantify their extent of breakdown based on
median and spread values; or use of the mixed Weibull distribution function
(Drechsler & Ferrua, 2016).
Oral digestion involves the enzymatic hydrolysis of starch, where salivary
α-amylases lightly hydrolyzes starch to generate fractions of maltose, maltotriose,
and α-limit dextrin (group of low-molecular-weight carbohydrates, 3–9 glucose
polymers) (Guyton & Hall, 2006). In addition, the mucin (glycoprotein) present in
the  saliva acts as a lubricant in bolus formation, contributing to the hydration of
the fragmented food (Bornhorst & Singh, 2012). In this first step of digestion, it is
to be expected that starch hydrolysis is limited due to the fact that granules remain
in the mouth for a short period. However, studies of oral digestion show the impor-
tance of the time/intensity of mastication, viscosity, the flow rate/concentration of
enzyme and pH on the salivary α-amylase activity on starch digestibility
(Butterworth, Warren, & Ellis, 2011). Moreover, in solid starchy foods that require
significant physical breakdown during digestion, breakdown behavior is affected by
the initial hardness and the rate of softening. The rate of softening is a function of
the food structure (e.g., porosity, density) and the amount of acid and enzyme secre-
tions that have entered the food matrix (e.g., hydration in physiological gastric con-
ditions) (Bornhorst, Ferrua, & Singh, 2015). Texture changes fit the Weibull model:

Ht
= e( )
− kt β
(4)
H0

Where Ht (N) is the hardness at time t (min), H0 is the initial hardness (N), k is the
scale parameter (min−1); t is the digestion time (min); and β is the distribution shape
factor (non-dimensional).
Although the salivary enzymatic degradation of starch has been considered
insignificant (not more than 5% of all starches) in comparison to pancreatic amylase
in the small intestine, it may influence the final digestive process by affecting starch
granules and the food structure (Hoebler et al., 1998). Individuals with high salivary
amylase concentrations may be better adapted to ingest starches, whereas those
individuals with low amylase activity may be at greater risk of insulin resistance and
diabetes if chronically ingesting starch-rich diets (Mandel & Breslin, 2012). The
Influence of Physical and Structural Aspects of Food on Starch Digestion 311

salivary flow rate and its composition can be stimulated by differences in starch-­
based foods. Engelen, Fontijn-Tekamp, and Van Der Bilt (2005) studied the influ-
ence of the characteristics of various starchy products on the swallowing threshold,
under the hypothesis that the urge to swallow food could be triggered by a threshold
level in both food particle size and lubrication of the food bolus. The influence of
oral physiology on the swallowing threshold was determined by measuring the sali-
vary flow rate, maximum bite force and masticatory performance. They used about
10 cm3 of bread, toast, Melba toast, breakfast cake, peanuts or cheese to determine
the influence on the swallowing threshold of various food characteristics (e.g., hard-
ness, moisture, and fat) and showed that salivary flow rates were significantly and
negatively correlated with the number of mastication (chewing) cycles of melba
toast and breakfast cake. Hence, subjects with more saliva needed less chewing
cycles for these dry foods.
Likewise, inactivation of salivary α-amylase may be influenced by low pH envi-
ronment or food composition, affecting the amount of hydrolyzed starch in the
mouth. Thus, the inhibition capacity to α-amylase by certain food components can
be used for its potential ability to modify the postprandial glycemic response. Some
α-amylase inhibitors are naturally present in cereals or legumes, or they can be
incorporated by formulation or food processing. Phenolic compounds, phytochemi-
cals in grains or crops, and bioactive compounds in seaweed have α-amylase inhibi-
tory properties. Heo et  al. (2009) indicated that diphlorethohydroxycarmalol
(DPHC) isolated from a brown alga might be a potent inhibitor for α-glucosidase
and α-amylase. The increase of postprandial blood glucose levels was significantly
suppressed in the DPHC-administered group compared to the streptozotocin-­
induced diabetic or normal mice.
After the food bolus is formed, it is swallowed and it travels through the esopha-
gus by muscle contractions (peristalsis) down to the stomach. The bolus transport
along the esophagus is influenced by the rheological behavior of starch-rich bolus
linked to viscosity, the swallowing rate, and the esophagus radius (Mackley et al.,
2013; Moritaka & Nakazawa, 2009).

1.3  Stomach Digestion

In the stomach, gastric secretions (acids, enzymes, and electrolytes) are secreted
from the stomach wall. With the help of weak peristaltic contraction waves, they
come into contact with the bolus to generate the chyme (semifluid mixture), and
they inactivate the salivary α-amylase due to their low pH. Not all amylase is inac-
tivated at the same time as the profile of pH inactivation is affected by chyme char-
acteristics (e.g., flow rate or viscosity), food pH and composition, or the amount of
bolus. The pH is adjusted to the characteristics of the bolus and its behavior during
its gastric passage. As such, the buffering capacity of the food probably prevents its
immediate and homogenous acidification in the stomach. At the same time, the gas-
tric pH may influence the rate of bolus hydration/disintegration, uptake of
312 I. Contardo and P. Bouchon

Fig. 5  Acidic water (AW) penetration front profiles as a function of distance (μm) from the carrot
center (solid circles) and Edam cheese (white circles) soaked for 10 min at pH 1.5 (a) and at pH
7.0 (b) [extracted from Van Wey et al. (2014)]

secretions, or food softening during digestion, promoting or limiting the release of


nutrients by the relationship between pH and the efficacy of different enzymes in the
gastric fluid. Van Wey et al. (2014) showed that the diffusion of acid water into a
food particle is pH dependent, and will depend upon the food structure as demon-
strated by the different penetration front profiles between a raw carrot core and
Edam cheese (see Fig. 5). Mennah-Govela and Bornhorst (2016) showed that the
uptake of gastric fluids into sweet potatoes during simulated gastric digestion modi-
fies the food macro- and micro-structures. The pH of the gastric juice increased
from 1.8 to 3.1 ± 0.5, and the pH of the food sample decreased (from 7 ± 0.3 to
3.7 ± 0.4) during 240 min of the gastric digestion period. The acidity of the sweet
potatoes cubes was significantly influenced by the cooking method and gastric
digestion time (p < 0.0001) as well as their interaction (p < 0.0001). In addition, the
mass transfer in sweet potatoes was influenced by cooking methods, demonstrating
the importance of the microstructure generated during processing on acid and water
diffusion into foods during digestion. Also, Bhattarai, Dhital, and Gidley (2016)
showed that increasing pH clearly results in more damage to starch, expressed in
terms of pitting (effect supported by SEM in the study) and consequently higher
hydrolysis. Wheat flour starch was hydrolyzed using various concentrations of
α-amylase ranging from 1 × 103 IU/mL (0.5 U/mg starch) to 12 × 103 IU/mL (6 U/
mg starch) at six different pH values (2–7).
Gastric digestion has physical–chemical implications for starchy products.
Diffusion of gastric fluids into the bolus and solid loss from the bolus to the gastric
site take place. Gastric mixing (muscle contractions), breakdown of the bolus into
small fractions, and interactions with gastric secretions damage starch granules,
leading to starch hydrolysis (Guyton & Hall, 2006). The flow rate of gastric secre-
tions and mixing are not homogenous during gastric digestion. These parameters
can be influenced by food structure, while starch granules that are not accessible
may act as physical barriers to the free diffusion of fluids into the bolus or disinte-
grated food particles. This modifies the contact rate between the active site and
enzyme action, and so it may  limit the rate and/or extent of starch hydrolysis.
Influence of Physical and Structural Aspects of Food on Starch Digestion 313

Fig. 6  Available glucose (mg/mL) during 180 min in vitro small intestinal digestion of control (no
fibre), locust bean gum, guar gum, fenugreek gum, xanthan gum, flaxseed gum, and soy soluble
polysaccharide-fortified solutions [extracted from Fabek et al. (2014)]

Therefore, the mass transfer in gastric starch digestion may be affected by viscosity,
food components, and gastric movements.
Viscosity of the bolus can affect the bioaccessibility of starch granules and the
quantity of glucose available for absorption, and it may also influence the glucose
concentration in the bloodstream. A study developed by AlHasawi et  al. (2017)
demonstrated that gastric viscosity differed considerably between commercially
available oat products, instant oats, steel cut oats, and oat bran, using TNO Intestinal
Model (TIM-1). Instant oat and oat bran viscosities were highest at the onset of
digestion and decreased over time, whereas the viscosity of steel cut oats at the
onset of digestion was the lowest level of viscosity observed, increasing over time,
and affecting the rate of starch digestion. Starch content was directly proportional to
total bioaccessible sugars and the rate of sugar release was slowest for steel cut oats
and quickest for oat bran. Thus, an increase in gastric viscosity may lead to reduc-
tion in the diffusion of hydrolyzed glucose. Likewise, cereals that are high in s­ oluble
fiber such as β-glucan may induce a high level of viscosity of the digesta once it has
reached the small intestine, where the absorption of glucose occurs. The viscosity of
soluble fibers depends on their ability to resist changes during gastrointestinal
digestion (Würsch & Pi-Sunyer, 1997). Fabek, Messerschmidt, Brulport, and Goff
(2014) demonstrated that dietary fiber influences the glucose diffusion of in vitro
small intestinal digestion in a simulated food model, which included protein and
starch (see Fig. 6). Villemejane, Wahl, Aymard, Denis, and Michin (2015) investi-
gated the effects of fiber in biscuit composition on the viscosity generated during
digestion using TIM-1 (stomach, duodenum, jejunum and ileum). The results
showed a significant effect of viscous soluble fibers on the chyme viscosity, up to
the ileal compartment. In the stomach, during the first hour of the digestion, fibers
were progressively liberated from the matrix and solubilized, which allowed main-
314 I. Contardo and P. Bouchon

tenance of the viscosity of the gastric content. These findings suggest that the inclu-
sion of more resistant to digestion ingredients, such as hydrocolloids (e.g., xanthan
gum, guar gum, or soy soluble polysaccharide) or soluble fibers (pectin, mucilages,
or β-glucan) in a food might be an effective strategy in lowering postprandial glyce-
mic responses in humans (Würsch & Pi-Sunyer, 1997).
The effect of viscosity on gastric emptying has shown diverse results. Increasing
bolus viscosity may delay gastric emptying rate, so it may increase satiety and mod-
ulate postprandial glycemic responses (Marciani et al., 2001; Zhu, Hsu, & Hollis,
2013). In contrast, Bornet et al. (1990) concluded that the α-amylase susceptibility
of test carbohydrates (25 g starch or equivalent glucose units) is a determining fac-
tor in the insulin response of healthy subjects, while viscosity of the test meals and
the gastric emptying rate have no effect.
The rate of bolus disintegration has also been shown to play a key role in gastric
emptying, as well as possibly influencing postprandial blood glucose levels. This is
conditioned by food composition or structure. Bornhorst and Singh (2013) demon-
strated that the disintegration rate and profile of bread boluses were significantly
influenced by bread composition in both static and shaking conditions. Each bread,
almond wheat, barley, rye, sourdough, wheat, or white, was characterized by its
moisture content, firmness, and water holding capacity. The initial moisture content
of breads also influenced the amount of gastric secretions absorbed. The total
amount of fluids absorbed by the bolus seemed to be inversely proportional to the
initial bread moisture content, while  the firmness of bread and its water holding
capacity were found to be complementary food properties that must be considered
to explain the differences detected in mass retention profiles. Additionally, the gas-
tric disintegration of a bolus can be affected by the properties of the gastric fluids
(composition and rheology) and the gastric movements (stomach motility and antral
contractions) (Kong & Singh, 2008). The structural breakdown of food has a signifi-
cant influence on starch hydrolysis, both in terms of bolus formation and disintegra-
tion (Bornhorst & Singh, 2012). Various studies have shown that different foods will
produce different particle size distributions in the bolus during oral processing. For
example, the bolus produced after mastication of bread demonstrated a bimodal
distribution of particle sizes (30  mm, 500  mm) (Hoebler, Devaux, Karinthi,
Belleville, & Barry, 2000). Particle sizes in a bolus have been linked to the rate of
in  vitro enzymatic degradation. Ranawana, Monro, Mishra, and Henry (2010)
reported that the degree of particle size, due to mastication, correlated with the rap-
idly available starch content (RDS) in chewed rice bolus. The quantity of undi-
gested material remaining at the end of the 120-min digestion correlated significantly
with the percentage of particles greater than 2000 μm in masticated rice.
After gastric digestion stomach emptying occurs, where the chyme is transported
from the stomach into the small intestine by intensive peristaltic contractions in the
antrum. The motion of the gastric fluids causes dramatic changes on the bolus struc-
ture, affecting both the glucose release and glucose diffusion during the digestion of
starch-based products. Kozu et al. (2010) investigated the effect of intra-gastric flow
on food digestion using computational fluid dynamics. They analyzed the flow phe-
nomena induced by gastric peristalsis with different fluid viscosities, focusing on
Influence of Physical and Structural Aspects of Food on Starch Digestion 315

the antral contraction wave (ACW), as the antrum is considered to act as a grinder


and mixer of  the swallowed foods and a pump for gastric emptying. The study
showed that the flow of the model gastric contents greatly changed in and close to
the region occluded by ACW. Thus, the gastric fluid motions induced by peristalsis
would promote the mixing of digestive enzymes with the gastric contents.
Furthermore, the introduction of mixing increases both glucose release and glucose
diffusion (Dhital, Dolan, Stokes, & Gidley, 2014).

1.4  Intestinal Absorption

The chyme is transported to the small intestine, where the starch is predominantly
hydrolyzed (~80%) to maltose, maltotriose, α-limit dextrins, or small glucose poly-
mers by pancreatic α-amylase from pancreatic secretions. These end-products are
further hydrolyzed to glucose by intestinal epithelial enzymes (Dhital, Warren,
Butterworth, Ellis, & Gidley, 2017). In addition, the intestinal motility process is
generated by three movement patterns: peristalsis, segmentation, and pendular
movements, which induce the mechanical digestion of starch. Peristalsis causes
propulsions that move the intestinal content in the anal direction, and segmentation
contractions cause mixing to promote absorption of nutrients and water. Finally, the
starch is absorbed in the form of glucose in the epithelium into the bloodstream,
through the brush border of the epithelial cells. In general terms, the absorption of
glucose can be represented by convection and diffusion processes across the intesti-
nal wall. Convective transport can be considered as a result of the flow induced by
the intestinal movements that transports and mixes chyme along the intestinal
lumen. Regarding the rate of movement of material from high to low concentration
by diffusion, it can be described mathematically by Fick’s laws of diffusion.
The small intestine can be divided into three parts: duodenum, jejunum, and
ileum. The entry of chyme into the upper portion of the duodenum triggers a key
event in the beginning of the intestinal phase of nutrient digestion. In this part, the
pancreatic secretions (enzymes, bicarbonate, and water) are secreted due to the
presence of chyme by pancreatic glands and epithelial cells. The characteristics of
pancreatic secretions related to viscosity, pH, and the flow rate/activity of enzymes
depend on the type and the amount of starch present in the chyme. Hence, again
when in vitro gastrointestinal methods are used to study starch digestion it is impor-
tant to understand mass transfer in the small intestine, as both biological and engi-
neering approaches are determining aspects to control and achieve realistic results
correlated with in  vivo results. Regarding the enzymatic activity of pancreatic
α-amylase, some molecules present in foods have been shown to inhibit α-amylase
in the intestine. Studies on humans have shown that natural α-amylase inhibitors
isolated from wheat significantly reduced glucose absorption and the peak of post-
prandial glucose in healthy and type 2 diabetic subjects (Lankisch, Layer, Rizza, &
DiMagno, 1998; Slavin, 2004). The plant forming part of the starch-based foods
have chemical constituents with the potential to inhibit α-amylase activity. For
316 I. Contardo and P. Bouchon

example, the chemical structures of flavonoids and polyphenols have been shown to
inhibit α-amylase activity and can reduce blood glucose levels after starchy foods
have been eaten. This could be potential constituents for controlling type 2 diabetes
(De Sales, De Souza, Simeoni, Magalhães, & Silveira, 2012; Lo Piparo et al., 2008;
Nyambe-Silavwe & Williamson, 2016). Likewise, the characteristics of intestinal
motility have proven to be relevant to modulate starch digestibility (Jaime-Fonseca,
Gouseti, Fryer, Wickham, & Bakalis, 2016). Gouseti et al. (2014), developed in vitro
intestinal models to study the effect of gut motility on the accessibility of glucose
from model solutions, using a range of food hydrocolloids (guar gum, carboxy-
methyl cellulose, pectin), and showing how mass transfer has an influence on nutri-
ent bioaccessibility. The study showed that the presence of gum guar and pectin
have a significant effect in retarding simulated glucose accessibility, and these
results appear to be more pronounced at viscosities levels of around 0.01  Pa·s.
Systems with lower viscosities showed enhanced mass transfer levels. The data
were analyzed using engineering principles and dimensionless numbers that charac-
terize the fluid flow (Reynolds number) and mass transfer (Sherwood number) in
the gut. The flow behavior can be determined by the velocity of the peristaltic flow
and the physical properties of chyme. Sherwood numbers represent the ratio of
convective to diffusive mass transfer coefficient. The absorption of the glucose
involved transportation from the lumen (chyme) to the dialysis membrane, passing
through the membrane, and transfer to the recipient fluid. This three-stage process
was characterized by the luminal mass transfer coefficient, (Klumen, m/s), diffusion
(described by coefficient Dmembrane, m2/s) through the membrane of thickness Zmembrane
(m), and the recipient side’s mass transfer coefficient (Krec, m/s). The following
equation offers the relationship between the local and overall transfer coefficient.

1 1 Z 1 1 1
= + membrane + = + (5)
K overall K lumen Dmembrane K rec K lumen K system

In addition, it should not be forgotten that emotional stimuli intrinsic to the indi-
vidual also influence the flow of gastrointestinal secretions, and can therefore affect
starch digestion in different ways.

2  I nfluence of Food Composition on the Variability of Starch


Digestion

The nutritional quality of starch is associated with its rate of digestion and glucose
absorption. Starch bioaccessibility and glucose release may differ depending on
starch structure and the form in which the food structure is disintegrated, some with
starch being rapidly and others slowly digested. Three aspects of the food composi-
tion are important to highlight related to variability of starch digestion: source of
starch, interaction of starch with other components, and the presence of dietary fiber.
Influence of Physical and Structural Aspects of Food on Starch Digestion 317

100
Rapidly Digestible Starch content, RDS [%] 90

80

70

60

50

40

30

20

10

0
–1 +1 –1 +1 –1 +1
Feed moisture Screw speed Temperature

High amylose maize starch Maize starch Waxy maize starch

Fig. 7  Rapidly digestible starch, RDS (%) of high amylose maize starch (67% amylose content),
maize starch (28% amylose content), and waxy maize starch (6% amylose content). The distance
between two grid lines represents the least significant difference (LSD) [extracted from Robin
et al. (2016)]

2.1  Source of Starch Granules

Significant differences in the rate of digestibility and metabolic responses of starch-­


based products have been associated with botanical sources of starch. The amylose
and lipid content of the granules varies, forming starches with high amylose or high
lipid content. An increase in amylose content has been correlated with slower
digestibility (Frei, Siddhuraju, & Becker, 2003; Robin, Heindel, Pineau, Srichuwong,
& Lehmann, 2016) (see Fig.  7). In addition, the amylose–lipid complex reduces
susceptibility to α-amylase. Thus, complexed amylose may help to decrease the
glycemic response of a food product (Hasjim, Ai, & Jane, 2013).
Wheat starch is a major dietary source and widely incorporated in processed
products such as snacks that have a high rate of consumption in the human diet. The
digestibility of white wheat bread is a typical example; in fact, this processed food
is used as a reference during glycemic responses assays in the same way as glucose
solutions (Englyst et al., 1999). Chemical characteristics related to amylose content
(waxy, normal, or high amylose) can regulate starch digestibility. Chen et al. (2016)
concluded that the gastrointestinal digestion rate of waxy wheat starch was higher
than that of normal wheat starch in the initial stages, and that a higher degree of
crystallisation limited the digestion rate and extent. Thus, the bioaccessibility of
wheat starch has critical implications for its digestion.
318 I. Contardo and P. Bouchon

The particle size of wheat grain is determinant forthe digestion rate of starch and
consequent glucose responses. Heaton, Marcus, Emmet, and Bolton (1988) demon-
strated that in  vitro starch hydrolysis by pancreatic α-amylase was faster with
decreasing particle size, and the peak of postprandial plasma glucose was greater
for fine-flour wheat than that for cracked or whole grains.
Regarding potato starch, this has a large granular size (<100 μm) and a concen-
tration of covalently bound phosphate in the amylopectin molecules as phosphate
monoesters and phospholipids (Singh, Singh, Kaur, Sodhi, & Gill, 2003). Some
studies have provided evidence that raw potato starch shows a reduced susceptibil-
ity to the action of amylase, due in part to its large granular size, based on the idea
that it is only the surface of the granule which is available for initial hydrolysis
(Cottrell, Duffus, Paterson, & Mackay, 1995); and the presence of phosphate groups
with a high B polymorph content, although these can be modified by processing and
storage. Warren, Zhang, Waltzer, Gidley, and Dhital (2015) demonstrated that native
potato starch was digested slowly and required more enzymes than maize to achieve
complete digestion. Potato starch granules can be completely digested in vitro given
enough enzyme and time, demonstrating the likely dependence of in vivo resistant
starch levels on endogenous enzyme activity and the small intestinal passage rate,
either or both of which may vary between foods and/or between individuals. García-­
Alonso and Goñi (2000) confirmed that boiled and mashed potatoes showed the
highest rate of digestion among raw flakes, oven-baked, French-fries, crisps, and
retrograded potato starch.
Legumes have acquired significant nutritional interest due to their rate of starch
digestion being lower in both in vitro and in vivo, compared to other starch sources
such as cereals. Their reduced bioavailability of starch has been attributed to the
presence of high levels of amylose (30–65%), a high content of viscous dietary fiber
components, the presence of antinutrients and B-type crystallites (Tharanathan &
Mahadevamma, 2003). These differences in their structural characteristics, as well
as their content of resistant starch, shows a slight reduction after processing in com-
parison to their raw form, which could allow them to be used as an alternative
source of resistant starch.

2.2  Interactions of Starch with Other Components

Most starch-based foods offer different metabolic responses even when they are
processed under similar conditions. These variances have been attributed to interac-
tions between starch with other food components such as proteins, lipids or sugars
during processing.
For instance, starch–gluten products such as bread are mainly processed by bak-
ing. In these types of product, gluten proteins (gliadins and glutenins) play a key
role in determining the baking quality of bread by conferring water absorption
capacity, cohesiveness, viscosity, and elasticity on dough (Wieser, 2007). During
processing, the proteins may physically become embedded in the starch, and inter-
Influence of Physical and Structural Aspects of Food on Starch Digestion 319

actions are a consequence of the attraction between positively and negatively


charged colloids due to the physical inclusion of starch in the gluten network
(Delcour et  al., 2000). Hence, starch–protein interactions may impair α-amylase
access and modify starch digestibility. The ability of starch–gluten interactions to
influence the bioaccessibility of starch depends on the characteristics of the protein
matrix and the degree of interaction. Previous studies have shown that in  vitro
digestibility of starch and its glycemic response increases when the starch–gluten
interaction is disrupted by sheeting passes or the mixing of dough (Kim et al., 2008;
Parada & Aguilera, 2011a, b).
Other important food components that can interact with starch during processing
are lipids, which mainly interact with the amylose molecules that affect the suscep-
tibility of starch to hydrolysis. The formation of amylose–lipid complex is given by
the ability of amylose to form inclusion complexes with polar lipids (e.g., mono-
glycerides, fatty acids, or lineal alcohol) during heat processing. The processing
temperature influences the type of complex formed and the time required for com-
plexation. Two types of complexes can be formed: (1) complexes with an amor-
phous structure (form type I) that melt at a lower temperature in a differential
scanning calorimetry (DSC) (10–30 °C) and (2) complexes with a crystalline struc-
ture (form type II) giving rise to the V-pattern in X-ray diffraction, although struc-
ture type I is not detectable using this technique. In addition, the formation of
amylose–lipid complex modifies the starch properties and functionality. Solubility
in water and swelling capacity are reduced, retrogradation is retarded, and suscepti-
bility to enzymatic hydrolysis is reduced (Parada & Santos, 2016). Starch digest-
ibility is reduced as complexed amylose becomes more resistant to digestive
enzymes than amylose, and decreases the swelling capacity of starch granules.
Consequently, there is less opportunity for enzymes to gain access to the granule
interior and less amylose leeches from the granules. Therefore, the rate and the
extent of hydrolysis of amylose–lipid complexes has been inversely related to the
degree of organization of helices into the aggregated structure, and complexes with
greater crystallinity are more resistant to enzymatic degradation (Seneviratne &
Biliaderis, 1991). Some authors have reported that the complex formation reduced
the digestibility of freshly gelatinized starch but increased the enzyme susceptibility
of stored starch, by competing with amylose retrogradation (Cui & Oates, 1999).
Furthermore, it is important to consider the time factor when making digestibility
analysis of lipid-rich products, which might affect the true results.
On the other hand, the presence of sugar influences the gelatinization by compe-
tition of available water in sugar–flour water systems, where sugar solubility may be
an important factor affecting gelatinization temperature in a limited water system
(Pareyt & Delcour, 2008). When the sugar content increases the gelatinization tem-
perature increases and the degree of starch gelatinization decreases (Hesso et al.,
2015). Accordingly, previous studies have demonstrated that the lower gelatiniza-
tion of starch involves a delay in starch digestibility.
Dietary fiber may also have an effect on starch digestibility. The effect of soluble
dietary fiber on starch digestibility is mainly attributed to increasing the bolus vis-
cosity once it has reached the small intestine, which is where the absorption of
320 I. Contardo and P. Bouchon

glucose occurs. This high viscosity delays glucose absorption. Some studies have
demonstrated that increased β-glucan intake improves glycemic control, indicating
it should be considered as a complementary mechanism in the treatment of patients
with type 2 diabetes (Jenkins, Jenkins, Zdravkovic, Würsch, & Vuksan, 2002). In
addition, Oh, Bae, and Lee (2014) established that under in vitro starch digestion,
decreasing levels of the inulin ratio in cakes resulted in a decrease in rapidly digest-
ible starch values. Interestingly, different types of soluble fiber have varying effects
on viscosity, and some studies have shown no correlation in all types of fiber
between high fiber content and reduced risk of diabetes, demonstrating that the
mechanics by which the hydrolysis of starch can be delayed are influenced by
intrinsic and extrinsic factors. Viscous fibers incorporated in food not only increase
the viscosity of the lumen but may also protect starch from enzymatic attack
(Gouseti et al., 2014). Furthermore, the presence of insoluble dietary fiber in com-
plex food systems has been associated with contributing to the control of diseases
such as obesity and diabetes, mainly due to the beneficial nutritional effects on
satiety and glycemic responses (Zhang & Hamaker, 2016). Thus, the compositional
differences link to white or whole-wheat flour influences the rate of starch digest-
ibility. Comparative studies between refined and whole grains (containing the outer
part of the bran) have demonstrated that whole grains of wheat with a high content
of dietary fibre helped lower the risk of diabetes mellitus (Liu et al., 2000).

3  Structural Aspects of Food and Starch Digestibility

Microstructural aspects of food can influence the digestion of starch. The bioacces-
sibility and bioavailability of starch is affected by the food matrix, influencing enzy-
matic functions and the residence time in the human stomach or intestine.
Accordingly, transformations of the food matrix and hormonal regulation mecha-
nisms can dominate the rate and extent of glucose release during gastrointestinal
transit (Parada & Aguilera, 2011a, b). Some microstructural aspects in solid starchy
foods and interrelated transformations are represented in Fig.  8, which involves
multiple reactions, mass transport, and glucose control mechanisms.
The main microstructural characteristics are linked to starch transformations,
particle size that can be obtained after the masticatory process, entrapment of starch
granules in the matrix, and viscosity and pH provided to the bolus by the compo-
nents forming the food matrix. Also, the physical texture (associated with the hard-
ness or density of the food) seems to have an impact on the availability of starch for
enzymatic digestion. As already discussed, the degree of starch gelatinization dras-
tically modifies the structure of materials in which the granules are entrapped, and
the digestion of starch and absorption of glucose within the digestive system.
Likewise, physical properties of the food bolus can be altered, providing a more/less
accessible structure, and so affecting the motility and susceptibility in the specific
activity of amylolytic enzymes. In addition, the hardness of a starchy food influ-
ences the particle size of its bolus. Chen, Khandelwal, Liu, and Funami (2013)
Influence of Physical and Structural Aspects of Food on Starch Digestion 321

Fig. 8  Microstructural aspects in starchy food and interrelated transformations involving multi-
ples reactions, mass transport, and glucose control mechanisms during digestion of starch [adapted
from Parada and Aguilera (2011a, b)]

studied the physical properties of food boluses, in particular the bolus particle size
distribution in relation to the hardness of the food. It was observed that bolus parti-
cle size decreased with the increase of food hardness (in cheese, peanuts, or cashew
nuts). The correlation between these two properties could be described by a power–
law relationship. Similarly, Alam et al. (2017) observed that the addition of 10% rye
bran had a significant effect on the structural, textural and mastication properties
both for puffs and flakes. The addition of rye bran increased hardness, decreased
crispiness, and increased the hydrolysis index of puffs and flakes to 89.7 and 94.5,
respectively, which was probably attributable to the increased number of particles in
the bolus. This was noticeable in the early phase of digestion, i.e. at 30 min, indicat-
ing that the disintegration process and consequently the particle size of the bolus
had an important role on the starch digestion rate. It is important to mention that
particle size is also influenced by inter-individual variability. Le Bleis, Chaunier,
Montigaud, and Della Valle (2016) proposed that bolus consistency could be
expressed by:

  Qs  (1+ n1)n 2 
n2


K = K 0 exp −α β   t  (6)
  W0  

This equation provided a basic model to describe the disruption of bread at different
stages of mastication, where K0 is the initial consistency index of the bread (Pa sn);
α is the plasticization coefficient (non-dimensional); Qs is the stimulated salivary
322 I. Contardo and P. Bouchon

flow (L min−1); t is the chewing time (min); w0, represents the median particle size
(mm); and the other coefficients [β, adjusted coefficient for salivation (non-­
dimensional); n1 and n2, adjusted exponents for fragmentation and salivation,
respectively (non-dimensional)] were obtained through fitting of experimental
results for breads enriched with fibers and examined in this study. The study showed
that bolus consistency decreased with chewing time, and this decrease was linked to
bolus moisture by a plasticization coefficient, which varied according to each indi-
vidual. Thus, the consistency of  the bread directly influenced bolus disruption
assessed by changes in viscosity.
Starch granules entrapped in the food matrix (e.g., plant cell, gluten network or
encapsulation), seem to be another mechanism that hinders the physical accessibil-
ity of starch and the diffusion of amylolytic enzymes in the starchy products.
Bhattarai, Dhital, Wu, and Gidley (2017) observed that the rate and extent of hydro-
lysis of starch and protein were greatly increased when the cell wall physical barrier
was removed by either mechanical or enzymatic processes. The authors used an
in-vitro dynamic model to observe that isolated legume cells have sufficient
mechanical strength to survive mixing conditions in a simulated rat stomach–duo-
denum. Also, the cell wall could limit digestibility by restricting starch gelatiniza-
tion during cooking, as water transfer (amount of liquid water molecules) into the
cell restricts the swelling of starch granules. Therefore, the use of whole grains (e.g.,
wheat, oat, barley, rye) in starchy products may result in low glycemic responses
due to the preservation of food particles in the gastrointestinal tract. This is because
hard solid foods are emptied more slowly from the stomach than soft foods.
Pasta products have shown slow and progressive starch breakdown and release of
sugar in the body, leading to low postprandial blood glucose and insulin responses
(Bjorck, Liljeberg, & Ostman, 2000). These wheat-based products vary in flour
variety, shape, type of drying, and proportion of protein added in their formulation,
promoting low glycemic responses. Pastas are prepared using durum wheat flour,
however it is also possible to incorporate dietary fiber ingredients and hydrocolloids
to increase their nutritional value. Sheeting, extrusion, drying, and cooking pro-
cesses confer the formation of different pasta structures by successive structural
changes of two main components, that is, starch and proteins, which provide a
potential to regulate the glycemic response of cereal foods. Thus, major structural
transformations occur during the cooking stage. Fardet, Hoebler, Bouchet, Gallant,
and Barry (1998) established that the presence of a structured and continuous pro-
tein network is an important factor in explaining the slow degradation of starch in
pasta. The authors proposed that the action of α-amylase may be limited at various
levels: (1) by the restricted accessibility and porosity of food structure; (2) by the
tortuosity of the protein matrix; (3) by the possible interactions of the α-amylase
with the protein matrix; and (4) by the structure of the starch granules in pasta;
demonstrating that the physical texture of starch-based food is a determinant factor
for bioavailability of starch in human digestion.
Furthermore, the modification of food structures with the addition of hydrocol-
loids (in order to modify rheological and textural aspects) may also have an effect.
Hydrocolloids influence the digestion and absorption of available carbohydrates in
Influence of Physical and Structural Aspects of Food on Starch Digestion 323

30000 100

Mass fraction of large particle (%)


Total peak area
Mass fraction of large particle
Total peak area (mV.S)

y = 9.7493x+12163
20000 R2=0.8144 80

10000 y = –0.019x+63.882 60
R2=0.9796

0 40
0 200 400 600 800 1000
Crust hardness (g)

Fig. 9  Correlation plots between total peak area, mass fraction of large particles, and crust hard-
ness [extracted from Gao et al. (2015)]

various ways. For instance, oat β-glucan in breads reduces the glycemic index (GI)
and glucose peak by 32–37% compared to a white wheat reference bread, and is
suitable for use in the baking of bread products (Ekström, Henningsson Bok, Sjöö,
& Östman, 2017). The addition of β-glucan into sugar cookies increased their attri-
bute of hardness, while affecting biscuit texture in turn decreased the carbohydrate
degradation and the rate of glucose absorption (Brennan, Samyue, & Abbot, 2004).
β-Glucan increased intestinal viscosity and delayed gastric emptying, which resulted
in a reduced rate of α-amylase action  and reduced intestinal nutrient uptake
(Thondre, Shafat, & Clegg, 2013). Likewise, the use of viscous soluble fiber as
Psyllium improves glycemic control in patients with type-2 diabetes mellitus
(Feinglos, Gibb, Ramsey, Surwit, & McRorie, 2013).
On the other hand, bread with various structures and textures provides different
chewing behavior and bolus characteristics, affecting the release of glucose. It has
been observed that there is an inverse relationship between food moisture content
and saliva secretion. Also, in the case of harder bread, the swallowing threshold of
particle size is smaller (see Fig. 9). The larger force and longer time for bread with
hard and dry crust during oral processing, resulted in turn in an extensive break-
down of the bread structure, which may contribute to the higher digestibility of
bread with a lower moisture content (Gao, Wong, Lim, Henry, & Zhou, 2015).

4  Process Design for the Control of Starch Digestion

Processed starchy foods are subjected to thermal processing to obtain desired


properties related to texture, quality, or nutritional aspects. Under processing,
the initial structure of the  food undergoes physical, mechanical and chemical
324 I. Contardo and P. Bouchon

transformations, affecting the form and structure of the final food product, and


the physical state of the starch granules that form part of it. Starch gelatinization
is promoted with the presence of liquid water and high temperature (>65  °C)
(Biliaderis, Maurice, & Vose, 1980; Eliasson, 1980). This phenomenon seems to
have an important influence on starchy food digestibility. Native starch shows
slow hydrolysis in contrast to partially or completely gelatinized starch, which
shows a faster rate of hydrolysis. This suggests that during processing native
granules lose their crystalline structure and become amorphous, thus facilitating
the action of amylolytic enzymes. Therefore, it is interesting to understand the
nutritional implications of gelatinization on starch digestibility, and how digestion
can be limited by processing conditions.

4.1  F
 ormation of Resistant Starch During Processing and Its
Relationship with Slow Starch Digestion

Resistant starch (RS) is a physiological concept that was initially defined as the
fraction of starch that was not hydrolyzed after 120 min of incubation with α-amylase
(Englyst et al., 1992; Sajilata, Singhal, & Kulkarni, 2006). However, it is now con-
sidered to be the fraction of starch and products of starch degradation that escape
digestion in the small intestine of healthy individuals. Five types of RS have been
established from RS1 (type I) to RS5 (type V) (Birt et al., 2013; Fuentes-Zaragoza,
Riquelme-Navarrete, Sánchez-Zapata, & Pérez-Álvarez, 2010).
In RS1, the starch granule is physically inaccessible to digestion due to its entrap-
ment in a matrix (e.g., grains, seeds, or food structure). In RS2, the resistance to
digestion is because the starch granule is in a granular form (e.g., compact structure
of granules such as ungelatinized resistant granules with type B- or C-polymorphism
of crystallinity). Ungelatinized granules are tightly packed in a radial pattern,
which limits the accessibility to digestive enzymes during hydrolysis. RS3, repre-
sents retrograded amylose formed during the cooling of gelatinized starch. It can be
formed when starch-based foods are thermally processed with enough water and
then cooled. Starch polymer chains begin to reassociate as double helices and can
form tightly packed structures stabilized by hydrogen bonding. In RS4, the resis-
tance to digestion is given by the formation of novel chemical bonds (e.g., cross-
linking with chemical agents). This type of RS includes chemically modified
starches. The last is RS5, which represents amylose–lipid complexes. Resistant
starches added to food matrices for health benefits are classified as functional fiber
by AACC (American Association of Cereal Chemists, 2001). In contrast to RS that
is naturally found in foods, it is considered dietary fiber. The dietary fiber definition
committee also reported that RS as a constituent of dietary fiber should be resistant
to digestion in humans and this should be assessed using methods that include
gelatinization steps to simulate cooking and processing (American Association of
Cereal Chemists, 2001).
Influence of Physical and Structural Aspects of Food on Starch Digestion 325

RS can be naturally found in significant levels in grains, seeds, legumes, tubers,


or unripe bananas. Some reasons for higher RS content are the crystallinity pattern
of the starch, the content of amylose, and ungelatinized starch. Tubers such as pota-
toes present a B-type pattern of crystallinity and legumes have a C-type pattern.
Bananas are a fruit consumed in a raw form, conserving a high content of ungelati-
nized starch (RS2) with a mixture of A-type and B-type patterns of crystallinity
depending on the varietal source (Zhang, Whistler, BeMiller, & Hamaker, 2005).
The B-type structure is hydrolyzed more slowly than the A-type by α-amylase,
β-amylase, and glucoamylase: this is probably due to surface area effects (Williamson
et al., 1992). Furthermore, the packing mode of the helices and water content are
different in the two polymorphs.
Resistant starches differ in their composition and structure, and the effects of
processing and storage on each type need to be analyzed separately. Different RS
can be generated during the processing of foodstuffs as such treatments modify the
normal starch granules and may promote RS formation or decrease its initial natural
content. For instance, thermal treatment (e.g., drying or extrusion) decreases the
presence of natural RS in products rich in starch, depending on their botanical
source. This is mainly because RS content decreases with increasing starch gelati-
nization, and therefore depends on the severity of the heat treatment and availability
of liquid water. RS formation may be increased due to starch retrogradation during
storage. In fact, keeping bread at room temperature for 3 days seems to be the best
way to further increase RS content [over 26%, as determined by Amaral, Guerreiro,
Gomes, and Cravo (2016)]. The water content in the dough seemed to influence the
extent of RS formation. The formed RS can be attributed to highly retrograded amy-
lose fraction (RS3). Tharanathan and Tharanathan (2001) isolated RS from wheat-­
based products: purified RS was a linear 1,4-linked α-d-glucan, which is derived
from the highly retrograded amylose fraction of starch. Numerous studies have
documented that RS provides benefits for health associated with decreasing levels
of glycemic response, which can modulate blood-glucose levels (Behall, Scholfield,
Hallfrisch, & Liljeberg-Elmstahl, 2006). Also, RS can be fermented by the colonic
microflora and produce short-chain fatty acids that provide the same physiological
response as functional fiber (Lattimer & Haub, 2010; Topping & Clifton, 2001).

4.2  Processing Conditions Over Starch Gelatinization

In the industry, starch-based products are dependent on the proper gelatinization of


starch to produce a desirable texture and mechanical properties. Native starch gran-
ules have ordered structures that are semi-crystalline and birefringent. During ther-
mal processing, starch granules suffer severe transformations by high temperature
and the presence of water. This order–disorder phase transition (gelatinization) is
associated with swelling of the granules (diffusion and water uptake by the amor-
phous zones), disruption of ordered structures (crystalline and molecular), and solu-
bilization of the micellar network (amylose leaching) (Lelievre & Liu, 1994). The
326 I. Contardo and P. Bouchon

extent of gelatinization determines the susceptibility of starch to enzymatic diges-


tion, as well as the extent of retrogradation (where the starch returns to the granular
state). Incomplete gelatinization permits low starch hydrolysis and slows glycemic
responses of starch-based products (e.g., bread, pasta, or potato chips) (Holm,
Lundquist, Björck, Eliasson, & Asp, 1988). Understanding how the degree of gela-
tinization can be limited might allow the starch digestion and consequent glucose
absorption to be modulated.
The most drastic effects on starch gelatinization are governed by temperature,
although there are also moisture-dependent interactions. In a complex and non-­
homogenous food system, physical barriers between granules and water molecules
can hinder heat transfer and water diffusion, influencing the kinetics of starch gela-
tinization. Water must pass through the resistance of the matrix (surface or internal),
hence gelatinization takes place in parts of the food where the water content and
temperature range are high enough (e.g., >30% (w/w) water content, for starch–
water systems). In baked foods (biscuits or bread) or pasta with limited water
through processing (flour to water ratio about 0.3–0.6), gelatinization occurs at low
levels of moisture content and when the range of gelatinization temperature is
extended (Schirmer, Zeller, Krause, Jekle, & Becker, 2014).
Likewise, most foods rich in starch are processed by means of boiling, baking,
extrusion, or frying, which can promote or limit water conditions, and in which the
starch can be partially or totally gelatinized. The low water content during process-
ing limits the degree of gelatinization. In dough matrices, there is competition
among the components for the available water, and the degree of starch ­gelatinization
will depend on the distribution of water in the system and the water activity on the
colloidal components. De la Hera, Rosell, and Gomez (2014) studied the impact of
dough hydration levels on in vitro starch hydrolysis of the rice flour used in gluten-­
free bread. The results indicated that the estimated glycemic index was higher in
breads with higher hydration (90–110% water content). Reduction of dough hydra-
tion limited starch gelatinization and hindered in  vitro starch digestibility. One
explanation is that increased  water content during the thermal process promotes
regions of amorphous starch, so making an attack by α-amylase more favorable
(Roder et al., 2009).
Furthermore, gelatinization is promoted under acidic environmental conditions,
although this could be conditioned by starch type and source. For example, Ohishi,
Kasai, Shimada, and Hatae (2007) showed that the absorption and gelatinization of
rice starch had been enhanced with the addition of acetic acid (0.2 M). An accelera-
tion of water absorption of starch by adding acetic acid promoted the hydration of
starch, leading to the enhancement of starch gelatinization. Also, the authors sug-
gested that the gelatinization process might be accelerated by the higher dissolution
and degradation of proteins under acidic conditions. This explanation agrees with
those studies using the Differential Scanning Calorimetry (DSC) technique, where
acid preferentially attacks the amorphous regions in the granule and the transforma-
tion of crystalline regions is changed, the crystallites becoming decoupled and no
longer destabilized by the amorphous parts. Consequently, starch crystallites of acid
modified starches melt at a higher temperature and the transition temperature range
Influence of Physical and Structural Aspects of Food on Starch Digestion 327

is broader. Furthermore, the addition of an alkali component has been found to sig-
nificantly enhance the swelling of starch granules and expedites the gelatinization
process (Wang et al., 2014).
Therefore, processing can offer alternatives for modifying the final nutritional
characteristics of foods. Also, limiting gelatinization of starch appears to be a suit-
able option to modulate starch hydrolysis and the glycemic responses of starch-­
based foods. Studies applied to both temperature and pressure have demonstrated
that the gelatinization temperature may be lowered by reducing the processing pres-
sure. Thus, the extent of swelling and granule disintegration, as well as leaching of
amylose, can be controlled. Various authors have shown that varying the pressure
conditions in traditional food processing such as boiling, drying, or frying, allow
some specific properties to be maintained such as color, antioxidant capacity, the
stability of specific compounds, or incomplete gelatinization. During the process of
high-pressure technology, the gelatinization of the starch granules occurs differ-
ently from damage by high temperatures, although by applying enough high pres-
sure it is possible to obtain complete starch gelatinization (Baks, Bruins, Janssen, &
Boom, 2008). In contrast, in low-pressure processing there is less damage of the
granules.

4.3  Low-Pressure Conditions Limiting Starch Digestibility

By means of pressure reduction during processing, it is possible to substantially


lower the boiling point of product moisture in a low-oxygen environment, this is the
main reason why vacuum technology is a recognized route to protect heat-sensitive
foods during dehydration. Applications may range from some traditional ones,
including vacuum evaporation in multiple effects, freeze-drying and vacuum dry-
ing, as well as some recent ones such as microwave vacuum drying and vacuum
frying (Dueik & Bouchon, 2011). Vacuum frying refers to the deep-fat frying pro-
cess that is carried out under pressures far below atmospheric pressure (Garayo &
Moreira, 2002). The processing conditions markedly decrease the boiling point of
water, reason why it corresponds to a vacuum dehydration process.
In order to compare vacuum and atmospheric frying, Mariscal and Bouchon
(2008) defined the concept of equivalent thermal driving force, which is achieved by
keeping a constant difference between oil temperature and the boiling point of water
at the working pressure, according to Eq. (7):

Thermal driving force = ∆T = Toil − Twater boiling point at working pressure (7)

Processing conditions under low pressure may affect the capacity of food building
blocks to develop an adequate structure during processing, to provide the required
quality attributes. This may be less relevant in raw materials that are already struc-
tured by nature, such as tubers, but still important. In formulated products, this may
be critical, since specific changes are needed to create a structure. In starchy
328 I. Contardo and P. Bouchon

Fig. 10  Semi-slab during Temperature


frying, showing the heat
fluxes and the temperature dd (t )/dt
profile (in red). The water
front at x = δ(t) moves at
Core Crust
dδ(t)/dt and is at Twater
boiling-point (function of P) Toil
[adapted from Bouchon
Twater boiling-point
and Pyle (2005)]

l
qcore qcrust qoil

x=0 x = d (t ) x = L /2

matrices, gelatinization is one of these critical steps, which requires the simultane-
ous presence of liquid water and temperature (above 55–60 °C). Frying is a complex
unit operation that involves simultaneous heat and mass transfer, resulting in
counter-­flows of water vapor (bubbles) and oil at the surface of the piece. After
immersion in the hot oil, the surface of the product is heated to the boiling point of
water and the crust begins to form. As frying progresses, the evaporation front,
which is at the boiling point of the interstitial liquid, will move towards the interior
(moving front), delimiting two very well-defined zones: the crust and the core
(Ziaiifar, Achir, Courtois, Trezzani, & Trystram, 2008). The crust is the result of
several alterations that mainly occur at the cellular and subcellular level, where the
temperature exceeds the boiling point of water (Bouchon & Aguilera, 2001). The
temperature within the core, on the other hand, cannot exceed the boiling point of
water, and thus holds liquid water. A diagram that reflects the aforementioned con-
ditions, showing the temperature profiles in each region (red lines), the heat fluxes
and the moving front, is presented in Fig. 10.
Accordingly, during vacuum frying, if the operating pressure defines a water
boiling point that is lower to the one required for starch gelatinization (e.g., <
55 °C), starch gelatinization may be impaired. In fact, the crust region will be able
to attain temperatures that are higher than the gelatinization temperature, but no
liquid water will be left for gelatinization to occur. Conversely, the core region will
have enough liquid water, but will be below the required temperature to induce
starch gelatinization. Ovalle, Cortés, and Bouchon (2013) demonstrated that when
the operational pressure was reduced up to 6.5 kPa, at a water boiling point of 38 °C,
no starch gelatinization was observed during heating in water and oil, in situ and in
real time, using vacuum hot-stage microscopy (see Fig. 11). This result was attrib-
Influence of Physical and Structural Aspects of Food on Starch Digestion 329

Fig. 11  Representation of vacuum hot-stage microscopy used for vacuum and atmospheric heat-
ing miniaturization [extracted from Ovalle et al. (2013)]

uted to the rapid evaporation of water before gelatinization was reached. In addition,
when the amount of water was reduced the gelatinization process occurred in a
broader range of temperatures.
Contardo, Parada, Leiva, and Bouchon (2016) assessed the effect of vacuum fry-
ing on starch gelatinization and in vitro digestibility of starch, in terms of the frac-
tions of rapidly available glucose (RAG), slowly available glucose (SAG), and
unavailable glucose (UG) fractions. The authors demonstrated that dough samples
in the form of sheets, made of wheat starch (88% d.b.), gluten (12% d.b.), and water,
and fried under vacuum (6.5 kPa, Twater-boiling-point = 38 °C), showed less starch gelati-
nization (28%), less rapidly available glucose (27%), and more unavailable glucose
(70%) than their atmospheric counterparts (which presented 99% starch gelatiniza-
tion, 40% rapidly available glucose, and 46% unavailable glucose, respectively),
and that the values were close to those of the raw dough. Recently, they comple-
mented their study, by assessing in vivo starch digestibility, after feeding Sprague-­
Dawley rats (Contardo, Villalón, & Bouchon, 2018). Results showed that
vacuum-fried dough had a maximal blood glucose level at 60 min, indicating a
slower glycemic response than that of samples fried under atmospheric counterparts
(maximal blood glucose level at 30 min), as shown in Fig. 12.
On the whole, both in vivo and in vitro studies were consistent and suggest that
starch digestibility can be altered through processing conditions by reducing the
operating pressure.
330 I. Contardo and P. Bouchon

195
Glucose
Blood glucose concentration (mg/dL)
180 AF (tep)
VF (tep)
165
Basal
150

135

120

105

90

75
0 15 30 45 60 75 90 105 120 135 150 165 180
Post-prandial time (min)

Fig. 12  In vivo starch digestibility expressed in terms of blood glucose concentration at various
postprandial times in rats given matrices with different degree of starch gelatinization, from
vacuum-­fried (VF, 9.9 kPa) and atmospheric-fried (AF) dough after frying up to bubble-end point,
as well as Glucose solution (1.2  g/Kg animal) as Control, and physiological serum as Basal
[extracted from Contardo et al. (2018)]

5  Conclusions

The food composition and the structure of processed starchy products, as discussed
in this chapter, may influence starch digestibility. The physical–chemical character-
istics of starch ingested (interactions with other components of the food matrix, as
well as transformations during processing) have a relevant impact on starch diges-
tion, and the associated glycemic response. Starch interactions with other compo-
nents may induce changes in the starch molecule (e.g., interactions between starch
and lipids), reducing starch digestibility. Also, limitations of free water availability
during processing can hinder the gelatinization process, precluding starch digest-
ibility. Overall, the principles highlighted here may help in the development of strat-
egies to modify starch-rich foods so as to reduce glycemic impact and improve the
impact of consuming such foods on health.

References

Alam, S. A., Pentikäinen, S., Närväinen, J., Holopainen-Mantila, U., Poutanen, K., & Sozer, N.
(2017). Effects of structural and textural properties of brittle cereal foams on mechanisms of
oral breakdown and in vitro starch digestibility. Food Research International, 96, 1–11.
AlHasawi, F. M., Fondaco, D., Ben-Elazar, K., Ben-Elazar, S., Fan, Y. Y., Corradini, M. G., et al.
(2017). In vitro measurements of luminal viscosity and glucose/maltose bioaccessibility for oat
bran, instant oats, and steel cut oats. Food Hydrocolloids, 70, 293–303.
Influence of Physical and Structural Aspects of Food on Starch Digestion 331

Amaral, O., Guerreiro, C.  S., Gomes, A., & Cravo, M. (2016). Resistant starch production in
wheat bread: Effect of ingredients, baking conditions and storage. European Food Research
and Technology, 242(10), 1747–1753.
American Association of Cereal Chemists. (2001). The definition of dietary fibre. Cereal Foods
World, 46(3), 112–126.
Baks, T., Bruins, M. E., Janssen, A. E. M., & Boom, R. M. (2008). Effect of pressure and tem-
perature on the gelatinization of starch at various starch concentrations. Biomacromolecules,
9(1), 296–304.
Behall, K. M., Scholfield, D., Hallfrisch, J., & Liljeberg-Elmstahl, H. (2006). Consumption of both
resistant starch and B-glucan improves postprandial plasma glucose and insulin in women.
Diabetes Care, 29(5), 976–981.
Bhattarai, R. R., Dhital, S., & Gidley, M. J. (2016). Interactions among macronutrients in wheat
flour determine their enzymic susceptibility. Food Hydrocolloids, 61, 415–425.
Bhattarai, R. R., Dhital, S., Wu, P., Chen, X., & Gidley, M. (2017). Digestion of isolated legume
cells in a stomach-duodenum model: Three mechanisms limit starch and protein hydrolysis.
Food & Function, 8, 2573–2582.
Biliaderis, C. G., Maurice, T. J., & Vose, J. R. (1980). Starch gelatinization phenomena studied by
differential scanning calorimetry. Journal of Food Science, 45(6), 1669–1674.
Birt, D. F., Boylston, T., Hendrich, S., Jane, J., Hollis, J., Li, L., et al. (2013). Resistant starch:
Promise for improving human health. American Society for Nutrition, 4, 587–601.
Bjorck, I., Granfeldt, Y., Liljeberg, H., Tovar, J., & Asp, N. (1994). Food properties affecting
the digestion and absorption of carbohydrates. American Journal of Clinical Nutrition, 59,
699S–705S.
Bjorck, I., Liljeberg, H., & Ostman, E. (2000). Low glycaemic-index foods. The British Journal of
Nutrition, 83(Suppl 1), S149–S155.
Bornet, F., Bizais, Y., Bruley Des Varannes, S., Pouliquen, B., Delort Laval, J., & Galmiche,
J. (1990). Gastric emptying rate controls plasma responses to starch in healthy humans. British
Journal of Nutrition, 63, 207–220.
Bornhorst, G. M., Ferrua, M. J., & Singh, R. P. (2015). A proposed food breakdown classifica-
tion system to predict food behavior during gastric digestion. Journal of Food Science, 80(5),
R924–R934.
Bornhorst, G. M., & Singh, R. P. (2012). Bolus formation and disintegration during digestion of
food carbohydrates. Comprehensive Reviews in Food Science and Food Safety, 11(2), 101–118.
Bornhorst, G. M., & Singh, R. P. (2013). Kinetics of in vitro bread bolus digestion with varying
oral and gastric digestion parameters. Food Biophysics, 8(1), 50–59.
Bornhorst, G. M., & Singh, R. P. (2014). Gastric digestion in vivo and in vitro: How the struc-
tural aspects of food influence the digestion process. Annual Review of Food Science and
Technology, 5, 111–132.
Bouchon, P., & Aguilera, J. M. (2001). Microstructural analysis of frying potatoes. International
Journal of Food Science and Technology, 36, 669–676.
Bouchon, P., & Pyle, D. L. (2005). Modelling oil absorption during post-frying cooling I: Model
development. Food and Bioproducts Processing, 83(4), 253–260 Retrieved from http://linkin-
ghub.elsevier.com/retrieve/pii/S0960308505704971
Brennan, C. S., Samyue, E., & Abbot, N. (2004). Evaluation of starch degradation and textural
characteristics of dietary fiber enriched biscuits. International Journal of Food Properties, 7(3),
647–657.
Butterworth, P. J., Warren, F. J., & Ellis, P. R. (2011). Human α-amylase and starch digestion: An
interesting marriage. Starch/Staerke, 63(7), 395–405.
Chen, J., Khandelwal, N., Liu, Z., & Funami, T. (2013). Influences of food hardness on the particle
size distribution of food boluses. Archives of Oral Biology, 58(3), 293–298.
Chen, P., Wang, K., Kuang, Q., Zhou, S., Wang, D., & Liu, X. (2016). Understanding how the
aggregation structure of starch affects its gastrointestinal digestion rate and extent. International
Journal of Biological Macromolecules, 87, 28–33.
332 I. Contardo and P. Bouchon

Contardo, I., Parada, J., Leiva, A., & Bouchon, P. (2016). The effect of vacuum frying on starch
gelatinization and its in vitro digestibility in starch-gluten matrices. Food Chemistry, 197, 353–
358. https://doi.org/10.1016/j.foodchem.2015.10.028
Contardo, I., Villalón, M., & Bouchon, P. (2018). In vivo study on the slow release of glucose in
vacuum fried matrices. Food Chemistry, 245, 432–438.
Cottrell, J. E., Duffus, C. M., Paterson, L., & Mackay, G. R. (1995). Properties of potato starch:
Effects of genotype and growing conditions. Phytochemistry, 40(4), 1057–1064.
Cui, R., & Oates, C. G. (1999). The effect of amylose-lipid complex formation on enzyme suscep-
tibility of sago starch. Food Chemistry, 65(4), 417–425.
De la Hera, E., Rosell, C. M., & Gomez, M. (2014). Effect of water content and flour particle size
on gluten-free bread quality and digestibility. Food Chemistry, 151, 526–531.
De Sales, P.  M., De Souza, P.  M., Simeoni, L.  A., Magalhães, P.  D. O., & Silveira, D. (2012).
α-Amylase inhibitors: A review of raw material and isolated compounds from plant source.
Journal of Pharmacy and Pharmaceutical Sciences, 15(1), 141–183.
Delcour, J. A., Vansteelandt, J., Hythier, M. C., Abécassis, J., Sindic, M., & Deroanne, C. (2000).
Fractionation and reconstitution experiments provide insight into the role of starch gelatiniza-
tion and pasting properties in pasta quality. Journal of Agricultural and Food Chemistry, 48(9),
3774–3778.
Dhital, S., Dolan, G., Stokes, J. R., & Gidley, M. J. (2014). Enzymatic hydrolysis of starch in the
presence of cereal soluble fibre polysaccharides. Food & Function, 5(3), 579–586.
Dhital, S., Warren, F. J., Butterworth, P. J., Ellis, P. R., & Gidley, M. J. (2017). Mechanisms of
starch digestion by α-amylase—Structural basis for kinetic properties. Critical Reviews in
Food Science and Nutrition, 57(5), 875–892.
Dona, A. C., Pages, G., Gilbert, R. G., & Kuchel, P. W. (2010). Digestion of starch: In vivo and
in vitro kinetic models used to characterise oligosaccharide or glucose release. Carbohydrate
Polymers, 80(3), 599–617.
Drechsler, K. C., & Ferrua, M. J. (2016). Modelling the breakdown mechanics of solid foods dur-
ing gastric digestion. Food Research International, 88, 181–190.
Dueik, V., & Bouchon, P. (2011). Vacuum frying as a route to produce novel snacks with desired
quality attributes according to new health trends. Journal of Food Science, 76(2), 188–195.
Ekström, L. M. N. K., Henningsson Bok, E. A. E., Sjöö, M. E., & Östman, E. M. (2017). Oat
β-glucan containing bread increases the glycaemic profile. Journal of Functional Foods, 32,
106–111.
Eliasson, A. (1980). Effect of water content on the gelatinization of wheat starch. Starch – Stärke,
32(8), 270–272.
Engelen, L., Fontijn-Tekamp, A., & Van Der Bilt, A. (2005). The influence of product and oral
characteristics on swallowing. Archives of Oral Biology, 50(8), 739–746.
Englyst, K. N., Englyst, H. N., Hudson, G. J., Cole, T. J., & Cummings, J. H. (1999). Rapidly
available glucose in foods: An in vitro measurement that reflects the glycemic response. The
American Journal of Clinical Nutrition, 69(3), 448–454.
Englyst, H.  N., Kingman, S.  M., & Cummings, J.  H. (1992). Classification and measurement
of nutritionally important starch fractions. European Journal of Clinical Nutrition, 46(2),
S33–S50.
Fabek, H., Messerschmidt, S., Brulport, V., & Goff, H. D. (2014). The effect of in vitro diges-
tive processes on the viscosity of dietary fibres and their influence on glucose diffusion. Food
Hydrocolloids, 35, 718–726.
Fardet, A., Hoebler, C., Bouchet, B., Gallant, D. J., & Barry, J. L. (1998). Involvement of the pro-
tein network in the in vitro degradation of starch from spaghetti and lasagne: A microscopic
and enzymic study. Journal of Cereal Science, 27, 133–145.
Feinglos, M. N., Gibb, R. D., Ramsey, D. L., Surwit, R. S., & McRorie, J. W. (2013). Psyllium
improves glycemic control in patients with type-2 diabetes mellitus. Bioactive Carbohydrates
and Dietary Fibre, 1(2), 156–161.
Influence of Physical and Structural Aspects of Food on Starch Digestion 333

Frei, M., Siddhuraju, P., & Becker, K. (2003). Studies on the in vitro starch digestibility and the
glycemic index of six different indigenous rice cultivars from the Philippines. Food Chemistry,
83(3), 395–402.
Fuentes-Zaragoza, E., Riquelme-Navarrete, M.  J., Sánchez-Zapata, E., & Pérez-Álvarez, J.  A.
(2010). Resistant starch as functional ingredient: A review. Food Research International, 43(4),
931–942.
Gao, J., Wong, J. X., Lim, J. C. S., Henry, J., & Zhou, W. (2015). Influence of bread structure on
human oral processing. Journal of Food Engineering, 167, 147–155.
Garayo, J., & Moreira, R. (2002). Vacuum frying of potato chips. Journal of Food Engineering,
55(2), 181–191.
García-Alonso, A., & Goñi, I. (2000). Effect of processing on potato starch: In vitro availability
and glycaemic index. Die Nahrung, 44(1), 19–22.
Goñi, I., Garcia-Alonso, A., & Saura-Calixto, F. (1997). A starch hydrolysis procedure to estimate
glycemic index. Nutrition Research, 17(3), 427–437.
Gouseti, O., Jaime-Fonseca, M. R., Fryer, P. J., Mills, C., Wickham, M. S. J., & Bakalis, S. (2014).
Hydrocolloids in human digestion: Dynamic in-vitro assessment of the effect of food formula-
tion on mass transfer. Food Hydrocolloids, 42(P3), 378–385.
Guyton, A., & Hall, J. (2006). Textbook of medical physiology. Physiology (11th ed.). Philadelphia,
PA: Elsevier Saunders.
Hasjim, J., Ai, Y., & Jane, J.  (2013). Novel applications of amylose-lipid complex as resistant
starch type 5. In Y.‐. C. Shi & C. C. Maningat (Eds.), Resistant starch (pp. 79–94). Hoboken,
NJ: Wiley.
Heaton, K. W., Marcus, S. N., Emmet, P. M., & Bolton, C. H. (1988). Particle-size of wheat, maize,
and oat test meals – Effects on plasma-glucose and insulin responses and on the rate of starch
digestion in vitro. American Journal of Clinical Nutrition, 47(4), 675–682.
Heo, S.  J., Hwang, J.  Y., Choi, J.  I., Han, J.  S., Kim, H.  J., & Jeon, Y.  J. (2009).
Diphlorethohydroxycarmalol isolated from Ishige okamurae, a brown algae, a potent
a-­glucosidase and a-amylase inhibitor, alleviates postprandial hyperglycemia in diabetic mice.
European Journal of Pharmacology, 615(1–3), 252–256.
Hesso, N., Loisel, C., Chevallier, S., Le-Bail, A., Queveau, D., Pontoire, B., et  al. (2015).
Monitoring cake baking by studying different ingredient interactions: From a model system to
a real system. Food Hydrocolloids, 51, 7–15.
Hoebler, C., Devaux, M., Karinthi, A., Belleville, C., & Barry, J. (2000). Particle size of solid food
after human mastication and in vitro simulation of oral breakdown. International Journal of
Food Sciences and Nutrition, 51, 353–366.
Hoebler, C., Karinthi, A., Devaux, M. F., Guillon, F., Gallant, D. J., Bouchet, B., et al. (1998).
Physical and chemical transformations of cereal food during oral digestion in human subjects.
British Journal of Nutrition, 80(5), 429–436.
Holm, J., Lundquist, J., Björck, I., Eliasson, A.-C., & Asp, N.-G. (1988). Degree in vitro, of starch
gelatinization, and metabolic response in rats. American Journal of Clinical Nutrition, 47,
1010–1016.
Jaime-Fonseca, M. R., Gouseti, O., Fryer, P. J., Wickham, M. S. J., & Bakalis, S. (2016). Digestion
of starch in a dynamic small intestinal model. European Journal of Nutrition, 55(8), 2377–2388.
Jenkins, A. L., Jenkins, D. J. A., Zdravkovic, U., Würsch, P., & Vuksan, V. (2002). Depression of
the glycemic index by high levels of beta-glucan fiber in two functional foods tested in type 2
diabetes. European Journal of Clinical Nutrition, 56(7), 622–628.
Kim, E., Petrie, J., Motoi, L., Morgenstern, M., Sutton, K., Mishra, S., et  al. (2008). Effect of
structural and physicochemical characteristics of the protein matrix in pasta on in vitro starch
digestibility. Food Biophysics, 3(2), 229–234.
Kong, F., & Singh, R. P. (2008). Disintegration of solid foods in human stomach. Journal of Food
Science R: Concise Reviews and Hypotheses in Food Science, 73(5), 67–80.
Kong, F., & Singh, R.  P. (2010). A human gastric simulator (HGS) to study food digestion in
human stomach. Journal of Food Science, 75(9), E627–E635.
334 I. Contardo and P. Bouchon

Kozu, H., Kobayashi, I., Nakajima, M., Uemura, K., Sato, S., & Ichikawa, S. (2010). Analysis
of flow phenomena in gastric contents induced by human gastric peristalsis using CFD. Food
Biophysics, 5(4), 330–336.
Lankisch, M., Layer, P., Rizza, R. A., & DiMagno, E. P. (1998). Acute postprandial gastrointestinal
and metabolic effects of wheat amylase inhibitor (WAI) in normal, obese, and diabetic humans.
Pancreas, 17(2), 176–181.
Lattimer, J. M., & Haub, M. D. (2010). Effects of dietary fiber and its components on metabolic
health. Nutrients, 2(12), 1266–1289.
Le Bleis, F., Chaunier, L., Montigaud, P., & Della Valle, G. (2016). Destructuration mechanisms of
bread enriched with fibers during mastication. Food Research International, 80, 1–11.
Lelievre, J., & Liu, H. (1994). A review of thermal analysis studies of starch gelatinization.
Thermochimica Acta, 246(2), 309–315.
Liu, S., Manson, J. E., Stampfer, M. J., Hu, F. B., Giovannucci, E., Colditz, G. A., et al. (2000).
A prospective study of whole-grain intake and risk of type 2 diabetes mellitus in US women.
American Journal of Public Health, 90(9), 1409–1415.
Lo Piparo, E., Scheib, H., Frei, N., Williamson, G., Grigorov, M., & Chou, C. J. (2008). Flavonoids
for controlling starch digestion: Structural requirements for inhibiting human alpha-amylase.
Journal of Medicinal Chemistry, 51(12), 3555–3561.
Mackley, M. R., Tock, C., Anthony, R., Butler, S. a., Chapman, G., & Vadillo, D. C. (2013). The
rheology and processing behavior of starch and gum-based dysphagia thickeners. Journal of
Rheology, 57(6), 1533.
Mandel, A., & Breslin, P. (2012). High endogenous salivary amylase activity is associated with
improved glycemic homeostasis following starch ingestion in adults 1–3. The Journal of
Nutrition, 142, 853–858.
Marciani, L., Gowland, P. a., Spiller, R.  C., Manoj, P., Moore, R.  J., Young, P., et  al. (2001).
Effect of meal viscosity and nutrients on satiety, intragastric dilution, and emptying assessed
by MRI. American Journal of Physiology. Gastrointestinal and Liver Physiology, 280(6),
G1227–G1233.
Mariscal, M., & Bouchon, P. (2008). Comparison between atmospheric and vacuum frying of
apple slices. Food Chemistry, 107(4), 1561–1569.
Mennah-Govela, Y. A., & Bornhorst, G. M. (2016). Mass transport processes in orange-fleshed
sweet potatoes leading to structural changes during in vitro gastric digestion. Journal of Food
Engineering, 191, 48–57.
Minekus, M., Alminger, M., Alvito, P., Ballance, S., Bohn, T., Bourlieu, C., et al. (2014). A stan-
dardised static in vitro digestion method suitable for food – An international consensus. Food
Function, 5(6), 1113–1124. Retrieved from http://xlink.rsc.org/?DOI=C3FO60702J
Moritaka, H., & Nakazawa, F. (2009). The rheological and swallowing properties of rice starch.
Food Science Research, 15(2), 133–140.
Nalin, T., Venema, K., Weinstein, D. A., de Souza, C. F. M., Perry, I. D. S., van Wandelen, M. T.
R., et al. (2015). In vitro digestion of starches in a dynamic gastrointestinal model: An innova-
tive study to optimize dietary management of patients with hepatic glycogen storage diseases.
Journal of Inherited Metabolic Disease, 38(3), 529–536.
Nyambe-Silavwe, H., & Williamson, G. (2016). Polyphenol- and fibre-rich dried fruits with green
tea attenuate starch-derived postprandial blood glucose and insulin: A randomised, controlled,
single-blind, cross-over intervention. British Journal of Nutrition, 116(3), 443–450.
Oh, I. K., Bae, I. Y., & Lee, H. G. (2014). In vitro starch digestion and cake quality: Impact of the
ratio of soluble and insoluble dietary fiber. International Journal of Biological Macromolecules,
63, 98–103.
Ohishi, K., Kasai, M., Shimada, A., & Hatae, K. (2007). Effects of acetic acid on the rice gela-
tinization and pasting properties of rice starch during cooking. Food Research International,
40(2), 224–231.
Ovalle, N., Cortés, P., & Bouchon, P. (2013). Understanding microstructural changes of starch dur-
ing atmospheric and vacuum heating in water and oil through online in situ vacuum hot-stage
microscopy. Innovative Food Science & Emerging Technologies, 17, 135–143.
Influence of Physical and Structural Aspects of Food on Starch Digestion 335

Parada, J., & Aguilera, J. M. (2011a). Microstructure, mechanical properties, and starch digest-
ibility of a cooked dough made with potato starch and wheat gluten. LWT – Food Science and
Technology, 44(8), 1739–1744.
Parada, J., & Aguilera, J. M. (2011b). Review: Starch matrices and the glycemic response. Food
Science and Technology International, 17(3), 187–204.
Parada, J., & Santos, J.  L. (2016). Interactions between starch, lipids, and proteins in foods:
Microstructure control for glycemic response modulation. Critical Reviews in Food Science
and Nutrition, 56(14), 2362–2369.
Pareyt, B., & Delcour, J. A. (2008). The role of wheat flour constituents, sugar, and fat in low mois-
ture cereal based products: A review on sugar-snap cookies. Critical Reviews in Food Science
and Nutrition, 48(November 2014), 824–839.
Ranawana, V., Monro, J. A., Mishra, S., & Henry, C. J. K. (2010). Degree of particle size break-
down during mastication may be a possible cause of interindividual glycemic variability.
Nutrition Research, 30(4), 246–254.
Robin, F., Heindel, C., Pineau, N., Srichuwong, S., & Lehmann, U. (2016). Effect of maize type
and extrusion-cooking conditions on starch digestibility profiles. International Journal of Food
Science & Technology, 51(6), 1319–1326.
Roder, N., Gerard, C., Verel, A., Bogracheva, T. Y., Hedley, C. L., Ellis, P. R., et al. (2009). Factors
affecting the action of α-amylase on wheat starch: Effects of water availability. An enzymic and
structural study. Food Chemistry, 113(2), 471–478.
Sajilata, M. G., Singhal, R. S., & Kulkarni, P. R. (2006). Resistant starch – A review. Comprehensive
Reviews in Food Science and Food Safety, 5, 1–17.
Schirmer, M., Zeller, J., Krause, D., Jekle, M., & Becker, T. (2014). In situ monitoring of starch
gelatinization with limited water content using confocal laser scanning microscopy. European
Food Research and Technology, 239(2), 247–257.
Seneviratne, H. D., & Biliaderis, C. G. (1991). Action of α-amylases on amylose-lipid complex
superstructures. Journal of Cereal Science, 13(2), 129–143.
Singh, N., Singh, J., Kaur, L., Sodhi, N., & Gill, B. (2003). Morphological, thermal and rheological
properties of starches from different botanical sources. Food Chemistry, 81, 219–231.
Slaughter, S. L., Butterworth, P. J., & Ellis, P. R. (2001). Mechanisms of the action of porcine pan-
creatic α-amylase on native and heat treated starches from various botanical sources. Starch –
Advances in Structure and Function, 1525, 110–115.
Slavin, J. (2004). Whole grains and human health. Nutrition Research Reviews, 17(1), 99–110.
Tharanathan, R. N., & Mahadevamma, S. (2003). Grain legumes – A boon to human nutrition.
Trends in Food Science and Technology, 14(12), 507–518.
Tharanathan, M., & Tharanathan, R. N. (2001). Resistant starch in wheat-based products: Isolation
and characterisation. Journal of Cereal Science, 34, 73–84.
Thondre, P.  S., Shafat, A., & Clegg, M.  E. (2013). Molecular weight of barley b-glucan influ-
ences energy expenditure, gastric emptying and glycaemic response in human subjects. British
Journal of Nutrition, 110, 2173–2179.
Topping, D. L., & Clifton, P. M. (2001). Short-chain fatty acids and human colonic function: Roles
of resistant starch and nonstarch polysaccharides. Physiological Reviews, 81(3), 1031–1064.
Van Wey, A. S., Cookson, A. L., Roy, N. C., McNabb, W. C., Soboleva, T. K., Wieliczko, R. J.,
et al. (2014). A mathematical model of the effect of pH and food matrix composition on fluid
transport into foods: An application in gastric digestion and cheese brining. Food Research
International, 57, 34–43.
Villemejane, C., Wahl, R., Aymard, P., Denis, S., & Michin, C. (2015). In vitro digestion of short-­
dough biscuits enriched in proteins and/or fibres, using a multi-compartmental and dynamic
system (1): Viscosity measurement and prediction. Food Chemistry, 182, 55–63.
Wang, S., Luo, H., Zhang, J., Zhang, Y., He, Z., & Wang, S. (2014). Alkali-induced changes in
functional properties and in vitro digestibility of wheat starch: The role of surface proteins and
lipids. Journal of Agricultural and Food Chemistry, 62(16), 3636–3643.
Wang, J. P., Zeng, A. W., Liu, Z., & Yuan, X. G. (2006). Kinetics of glucoamylase hydrolysis of
corn starch. Journal of Chemical Technology and Biotechnology, 81(4), 727–729.
336 I. Contardo and P. Bouchon

Warren, F. J., Zhang, B., Waltzer, G., Gidley, M. J., & Dhital, S. (2015). The interplay of alpha-­
amylase and amyloglucosidase activities on the digestion of starch in in vitro enzymic systems.
Carbohydrate Polymers, 117, 192–200.
Wickham, M. J. S., Faulks, R. M., Mann, J., & Mandalari, G. (2012). The design, operation, and
application of a dynamic gastric model. Dissolution Technologies, 19(3), 15–22.
Wieser, H. (2007). Chemistry of gluten proteins. Food Microbiology, 24(2), 115–119.
Williamson, G., Belshaw, N.  J., Self, D.  J., Noel, T.  R., Ring, S.  G., Cairns, P., et  al. (1992).
Hydrolysis of A- and B-type crystalline polymorphs of starch by a-amylase, b-amylase and
glucoamylase 1. Carbohydrate Polymers, 18(3), 179–187.
Würsch, P., & Pi-Sunyer, X. (1997). The role of viscous soluble fiber in the metabolic control of
diabetes. Diabetes Care, 20(11), 1774–1780.
Zhang, G., & Hamaker, B. R. (2016). The nutritional property of endosperm starch and its con-
tribution to the health benefits of whole grain foods. Critical Reviews in Food Science and
Nutrition, 57, 3807–3817.
Zhang, P., Whistler, R., BeMiller, J., & Hamaker, B. (2005). Banana starch: Production, physi-
cochemical properties, and digestibility – A review. Carbohydrate Polymers, 59(4), 443–458.
Zhu, Y., Hsu, W. H., & Hollis, J. H. (2013). The impact of food viscosity on eating rate, subjective
appetite, glycemic response and gastric emptying rate. PLoS One, 8(6), 6–11.
Ziaiifar, A. M., Achir, N., Courtois, F., Trezzani, I., & Trystram, G. (2008). Review of mechanisms,
conditions, and factors involved in the oil uptake phenomenon during the deep-fat frying pro-
cess. International Journal of Food Science and Technology, 43(8), 1410–1423.
Zimmet, P., Alberti, K. G. M. M., & Shaw, J. (2001). Global and societal implications of the dia-
betes epidemic. Nature, 414(December 2001), 782–787.

You might also like